Fact-checked by Grok 2 weeks ago

Capillary number

The Capillary number () is a in that quantifies the ratio of viscous forces to interfacial tension (or capillary) forces acting on an between two immiscible s. It is defined by the \mathrm{Ca} = \frac{\mu [v](/page/V.)}{[\sigma](/page/Sigma)}, where \mu is the dynamic of the , v is the of the , and \sigma is the interfacial tension between the fluids. This parameter, first introduced as a correlating group in studies in 1927 and further developed by researchers like Leverett in the late 1930s and early 1940s, helps predict the behavior of fluid interfaces under dynamic conditions. In practical applications, the Capillary number is particularly significant in two-phase flows through porous media, where low values (typically \mathrm{Ca} \approx 10^{-6} to $10^{-5} during conventional waterflooding) indicate dominance of capillary forces, leading to trapping of residual saturation. Increasing to around $10^{-4} or higher—often through (EOR) techniques like flooding that reduce \sigma from ~30 mN/m to as low as $10^{-3} mN/m—mobilizes trapped non-wetting phases by shifting the force balance toward viscous effects. Beyond , governs phenomena such as droplet deformation in microfluidic devices, in porous materials, and the dynamic in capillary flows, with over 40 variant definitions existing to account for microscopic, macroscopic, or mixed scales depending on the context. When ≪ 1, surface tension stabilizes interfaces, while ≫ 1 signifies viscous dominance, enabling stripping or deformation mechanisms.

Fundamentals

Definition

The capillary number, denoted as Ca, is a dimensionless quantity in fluid mechanics that characterizes the ratio of viscous forces to interfacial tension (or surface tension) forces acting across a fluid interface during flow. Viscous forces arise from shear stress within the fluid, which resists relative motion between fluid layers, while surface tension represents the interfacial energy per unit area that minimizes the surface area of the interface. This ratio provides a measure of the relative importance of these competing forces in processes involving fluid interfaces, such as those in multiphase flows. The concept of the capillary number originates from dimensional analysis techniques in capillarity studies, formalized through the Buckingham π theorem introduced by Edgar Buckingham in 1914, which identifies key dimensionless groups governing physical phenomena like capillary action. Although the specific term "capillary number" gained prominence later in the context of petroleum engineering and multiphase flow research, it builds directly on these foundational methods for scaling fluid behaviors involving viscosity and surface tension. Physically, a low capillary number (Ca ≪ 1) indicates dominance of forces, leading to minimal deformation of fluid , such as the formation and maintenance of nearly spherical droplets or bubbles. Conversely, a high capillary number (Ca ≫ 1) signifies that viscous forces prevail, causing significant elongation or deformation of , as seen when droplets stretch under in flowing systems. This transition highlights the capillary number's role in predicting stability and flow regimes without requiring detailed geometric specifications.

Single-Fluid Formulation

The single-fluid formulation of the capillary number arises from a force balance between viscous forces driving the flow and forces resisting deformation at the fluid interface. In this context, the viscous force scales as F_v \sim \mu v L, where the \mu v / L acts over a characteristic area L^2, while the force scales as F_c \sim \sigma L, representing the energy required to extend the interface over a L. The ratio of these forces yields the capillary number \mathrm{Ca} = \frac{\mu v}{\sigma}, as the L cancels out, providing a dimensionless measure of viscous dominance over interfacial tension. This formulation can also be obtained through dimensional analysis using the Buckingham π theorem, which identifies dimensionless groups from the relevant physical variables: dynamic viscosity \mu, characteristic velocity v, and interfacial tension \sigma. With three variables and two fundamental dimensions (mass and time), the theorem predicts one dimensionless π-group, which takes the form \mathrm{Ca} = \frac{\mu v}{\sigma}. In the formula \mathrm{Ca} = \frac{\mu v}{\sigma}, the parameter \mu represents the fluid's resistance to shear deformation, quantifying internal during ; v is the characteristic speed, such as the average velocity in a conduit; \sigma denotes the interfacial , the work per unit area needed to expand the fluid-fluid ; and the is implicit in the scaling but absent in the final expression. The dimensionless nature of \mathrm{Ca} is confirmed by SI unit analysis: [\mu] = \mathrm{kg \cdot m^{-1} \cdot s^{-1}}, = \mathrm{m \cdot s^{-1}}, so [\mu v] = \mathrm{kg \cdot s^{-2}}; [\sigma] = \mathrm{N \cdot m^{-1}} = \mathrm{kg \cdot s^{-2}}; thus, [\mathrm{Ca}] = 1. For example, consider water (\mu = 1 \times 10^{-3} \, \mathrm{Pa \cdot s}, \sigma = 0.072 \, \mathrm{N/m} with air at 20°C) flowing at a low speed of v = 0.1 \, \mathrm{m/s} in a tube, as might occur in a simple capillary experiment; this yields \mathrm{Ca} \approx 1.4 \times 10^{-3}, indicating surface tension still plays a significant role.

Advanced Formulations

Multiphase Formulation

In multiphase flows, particularly immiscible displacements like waterflooding in oil reservoirs, the capillary number is adapted to account for interfacial tension between phases and the properties of the displacing fluid. The formulation is given by \text{Ca} = \frac{\mu_w v}{\sigma}, where \mu_w is the of the (typically the displacing fluid, such as ), v is the Darcy velocity, and \sigma is the interfacial tension between the and non-wetting (e.g., oil-water). This differs from the single-fluid case by emphasizing the ratio of viscous forces in the displacing to the interfacial forces resisting meniscus movement at the fluid-fluid . In porous media, influences the effective flow behavior, leading to modifications of the capillary number to incorporate phase-specific conductance. An effective capillary number can be defined as \text{Ca}_\text{eff} = \text{Ca} \sqrt{\frac{k_{rw}}{k_{rnw}}}, where k_{rw} and k_{rnw} are the relative permeabilities of the and non-wetting phases, respectively; this adjustment accounts for saturation-dependent reductions in permeability for each phase during displacement. The multiphase capillary number derives from combining for each phase with the across the . for the wetting phase states v_w = -\frac{k k_{rw}}{\mu_w} \nabla p_w, while the is p_c = \frac{\sigma \cos \theta}{r}, where \theta is the and r is the characteristic radius. At the displacement front, the viscous \nabla p \approx \frac{\mu_w v}{k k_{rw}} competes with the \nabla p_c \approx \frac{p_c}{L} (with L the ), yielding the capillary number as the ratio that determines whether viscous forces destabilize the , potentially causing viscous . In porous media, displacement fronts remain stable for \text{Ca} < 10^{-5}, where capillary forces dominate and lead to compact invasion; above this threshold, instabilities like viscous fingering emerge as viscous forces prevail. For typical reservoir conditions—\mu_w = 1 cP, v = 1 ft/day (approximately $3.5 \times 10^{-6} m/s), and \sigma = 30 dyn/cm (0.03 N/m)—the capillary number is on the order of $10^{-7}, indicating capillary-dominated flow with stable fronts under standard waterflooding.

Variations in Different Contexts

In non-Newtonian fluids, particularly those following a power-law rheology, the capillary number is defined using an effective viscosity derived from the power-law model, \mu_\text{eff} = K (\dot{\gamma})^{n-1}, where K is the consistency index (Pa·s^n), \dot{\gamma} is the characteristic shear rate (typically \dot{\gamma} \propto v / L with L a characteristic length), n is the flow behavior index (with n < 1 for shear-thinning and n > 1 for shear-thickening fluids), leading to \text{Ca} = \frac{\mu_\text{eff} v}{\sigma}. This adaptation accounts for the nonlinear relationship between and , enabling accurate prediction of flow regimes where rheological properties dominate capillary dynamics, such as in processing or biological fluids. In electrokinetics and magnetohydrodynamics, the capillary number is augmented to incorporate field-induced stresses on fluid interfaces. For electrokinetic flows involving charged interfaces, an \text{Ca}_e = \frac{\epsilon E^2 L}{\sigma} is employed, where \epsilon is the electric permittivity, E the applied strength, L the , and \sigma the surface tension; this quantifies the competition between Maxwell stresses and capillary forces in applications like electro-osmotic pumping. Analogously, in contexts, a \text{Ca}_m = \frac{\mu_0 \chi H^2 L}{\sigma} arises, with \mu_0 the , \chi the , and H the strength, capturing magnetic forces that deform interfaces in ferrofluids or flows. These are essential for systems where external fields modify interfacial stability beyond purely hydrodynamic effects. For high-speed flows where inertia cannot be neglected, the capillary number transitions toward dominance, with the (We = \rho v^2 L / \sigma) characterizing the ratio of inertial to capillary forces. This modification is particularly relevant in or jet breakup scenarios, bridging low- and high-speed regimes without fully abandoning capillary considerations. In , capillary numbers can reach values sufficient to observe phenomena like droplet formation or dynamics in devices, necessitating formulations that account for dimensions and conditions.

Physical Significance

Role in Multiphase Flows

In multiphase flows through porous media, the (Ca) governs the balance between viscous and capillary forces, dictating the dominant mechanisms and . At low Ca values, typically below 10^{-4}, capillary forces dominate, promoting capillary-dominated where the invading fluid advances along stable, compact pathways influenced by local wettability and geometry, resulting in relatively uniform progression without extensive ramification. This is characterized by the prevalence of interfacial , leading to controlled invasion patterns that minimize bypassing of the displaced . In such capillary-dominated conditions (typically Ca < 10^{-5}), snap-off—where the wetting swells in throats to pinch off non-wetting ganglia—becomes a key process, predominating in constricted pores and generating dispersed oil-in-water or water-in-oil emulsions that enhance mobility but can also increase pressure gradients due to droplet interactions. As increases to above 10^{-2}, viscous forces take precedence, shifting the behavior to viscous fingering, which produces unstable, branched, and dendritic patterns, especially when the displacing fluid has lower than the displaced one. This arises from perturbations at the that amplify under high rates, causing the invading to penetrate preferentially and leave behind significant trapped volumes of the resident fluid. In immiscible displacements, such as displacing , directly impacts residual saturation by controlling the mobilization of trapped ganglia; elevating reduces residual saturation as viscous stresses deform and dislodge blobs otherwise immobilized by capillary forces. The critical for initial mobilization lies between 10^{-6} and 10^{-5}, marking the onset where trapped phases begin to , with further increases yielding substantial desaturation. At intermediate Ca values, around 10^{-5} to 10^{-3}, coalescence of droplets becomes a key process that facilitates formation. Coalescence events, driven by thin film drainage between approaching interfaces, further stabilize or destabilize these emulsions, altering the effective of the multiphase system. Experimental core flooding tests confirm these transitions, revealing a Ca of approximately 10^{-6} to 10^{-5} where shifts from piston-like fronts—smooth, uniform advances with minimal —to shocked fronts characterized by abrupt changes and reduced capillary end effects. Below this range, capillary smooths the front, promoting stable but inefficient sweeping; above it, viscous dominance sharpens the profile, improving sweep efficiency in favorable scenarios.

Relation to Other Dimensionless Numbers

The capillary number () is closely related to the ( = ρ v L / μ), which quantifies the ratio of inertial forces to viscous forces in fluid flows. High values of generally correspond to low regimes, where viscous drag dominates over inertial effects at fluid interfaces, such as in slow, creeping flows through capillaries or porous structures. This distinction highlights the transition from inertia-driven to viscosity-driven interfacial deformation; for instance, in multiphase systems, low ensures that capillary and viscous forces govern meniscus shapes without turbulent disruptions. The combined parameter × encapsulates the full force balance among viscous, inertial, and surface tension effects, aiding in the analysis of free-surface flows where both momentum and interfacial stability are critical. A key interconnection exists between and the (We = ρ v² L / σ), which measures inertial forces relative to . Mathematically, We = × , revealing that the scales the product of viscous-to-capillary and inertial-to-viscous ratios, thereby marking the shift from capillary-dominated regimes (low We) to those where inertial forces disrupt interfaces (high We). This relation is particularly useful in predicting phenomena like droplet breakup or jet atomization, where increasing velocity elevates We beyond unity, overriding stabilization. In processing diagrams for complex fluids, trajectories governed by the Ohnesorge number (Oh = \sqrt{\frac{\mathrm{Ca}}{\mathrm{Re}}}) further illustrate how and jointly influence We-dependent outcomes. The Bond number (Bo = Δρ g L² / σ) complements by contrasting gravitational forces with forces, providing insight into buoyancy-driven versus viscous-driven multiphase dynamics. Low paired with high favors gravity drainage in vertical flows, as overcomes retention, enabling efficient phase segregation without significant viscous resistance. Conversely, high and low emphasize viscous displacement over , altering profiles in heterogeneous media. This comparison is essential for understanding flow stability and recovery efficiency in systems where body forces compete with interfacial tension. In porous media applications, the capillary number is frequently expressed as the group N_ca = v μ / σ, which integrates with and to enable in scaled experiments. This formulation allows researchers to match viscous-to-capillary ratios (N_ca) alongside inertial () and gravitational () effects, ensuring that laboratory models replicate field-scale multiphase transport behaviors, such as or . For example, experiments maintaining N_ca ≈ 10^{-5} to 10^{-4}, low (<1), and moderate (≈10^{-4}) isolate capillary dominance while validating predictions against gravitational or inertial perturbations. Such combined dimensionless groups underpin quantitative assessments of and residual saturations in subsurface flows.

Applications

In Enhanced Oil Recovery

In (EOR), the capillary number plays a pivotal role in waterflooding processes, where additives are employed to increase the displacing fluid's , thereby elevating the capillary number and facilitating the mobilization of residual oil trapped by capillary forces. Typically, conventional waterflooding operates at low capillary numbers on the order of 10^{-6} to 10^{-5}, leaving significant residual oil saturation, but addition can raise this value to 10^{-5} or higher by enhancing viscous forces relative to interfacial tension. This increase often results in additional oil recovery of 10-20% of the original , as demonstrated in laboratory corefloods and field pilots, by reducing oil trapping in throats and improving sweep efficiency. Chemical EOR techniques further leverage the capillary number to target residual , with flooding primarily lowering interfacial to dramatically increase the capillary number and promote detachment from rock surfaces. By reducing interfacial from millinewtons per meter to ultralow values (e.g., 10^{-3} mN/m), can elevate the capillary number by orders of magnitude, enabling the of otherwise immobilized at low capillary numbers. In alkaline flooding, the involves both interfacial reduction through in-situ generation and wettability alteration toward more water-wet conditions, which indirectly enhances the effective capillary number and alters its interpretation in . These combined effects in alkali--polymer systems can achieve capillary numbers sufficient to lower residual below waterflood levels, improving overall in heterogeneous reservoirs. The number is integral to simulation and scaling in EOR, particularly within fractional flow theory as embodied in the Buckley-Leverett equation, where it influences the and propagation of displacement fronts by balancing viscous and forces. Lower numbers stabilize the front against , leading to more uniform sweep and predictable recovery profiles in numerical models. For lab-to-field upscaling, matching numbers ensures that core-scale experiments translate to conditions, using desaturation curves to correlate residual oil saturation with local number variations and validate simulations across scales. A historical milestone in applying the capillary number to EOR occurred during 1970s field tests, such as polymer floods in reservoirs like the Minnelusa Formation, where elevated capillary numbers demonstrated miscible-like recovery efficiencies in otherwise immiscible systems. These pilots, including early implementations at fields like Skull Creek, highlighted how polymer flooding could overcome capillary trapping through improved mobility control, paving the way for widespread adoption of chemical EOR methods.

In Microfluidics and Lab-on-a-Chip

In microfluidics and lab-on-a-chip devices, the capillary number (Ca) plays a pivotal role in controlling droplet formation regimes, enabling precise manipulation of fluid interfaces at microscales. At low Ca values around 10^{-3}, capillary forces dominate, leading to the dripping regime where uniform, monodisperse droplets are generated through interfacial instabilities in geometries like flow-focusing or T-junctions. This regime is ideal for applications requiring discrete, size-controlled emulsions, as the balance favors surface tension over viscous shear. In contrast, at higher Ca values approaching 1, viscous forces prevail, transitioning to the jetting regime that produces elongated threads for continuous droplet streams, often used in high-throughput particle synthesis. The capillary number also governs passive versus active flow control strategies in these systems, extending principles like Washburn's law for self-propelled in capillaries. In passive setups, low Ca (typically 10^{-4} to ) enables pump-free operation, where drives fluid according to L(t) ∝ √t, facilitating autonomous mixing and transport in open or closed channels without external actuation. This contrasts with active control, where higher Ca allows syringe-driven flows for dynamic adjustments, but passive designs dominate for simplicity and portability. Such Ca ranges ensure efficient diffusive mixing in low-Reynolds environments, avoiding the need for mechanical pumps. In biomedical applications, Ca optimization enhances emulsion stability for drug delivery, where low Ca values promote robust, uniform droplets that encapsulate therapeutics for controlled, targeted release, minimizing coalescence and improving bioavailability. For instance, in (PDMS) channels fabricated via , tuning Ca to around 10^{-3} to 10^{-2} enables precise droplet encapsulation during , allowing isolation and study of individual cells in heterogeneous samples like , with applications in and diagnostics. Post-2020 advancements have integrated considerations with techniques, where the number scales inversely with channel (h/w), optimizing capillary filling rates for portable diagnostics. This enables compact, instrument-free devices, such as those for antigen detection using capillary-driven assays in PDMS structures, achieving rapid results with minimal sample volumes and enhancing point-of-care accessibility in resource-limited settings.

References

  1. [1]
    A Brief Review of Capillary Number and its Use in Capillary ...
    Jul 21, 2022 · Capillary number, understood as the ratio of viscous force to capillary force, is one of the most important parameters in enhanced oil recovery (EOR).
  2. [2]
    Capillary Number | Fundamentals of Fluid Flow in Porous Media
    The capillary number is a dimensionless number describing the relative importance of viscous forces to capillary forces during immiscible displacement.
  3. [3]
    capillary number - Energy Glossary - SLB
    A capillary number is a dimensionless group that characterizes the ratio of viscous forces to surface or interfacial tension forces. It is calculated as (μU) / ...
  4. [4]
    Capillary Number | Wolfram Formula Repository
    The capillary number represents the relative effect of viscous forces versus surface tension acting across an interface between a liquid and a gas.
  5. [5]
    Illustrations of the Use of Dimensional Equations | Phys. Rev.
    On physically similar systems; illustrations of the use of dimensional equations. E. Buckingham Bureau of Standards.Missing: Edgar capillary
  6. [6]
    Droplet dynamics in three-way microchannel: Breakup, sorting, and ...
    Aug 20, 2025 · A lower capillary number indicates that interfacial tension dominates viscous stress, causing droplets to retain an almost spherical shape ...
  7. [7]
    Capillary Number (Ca) – CHEM-ENG 535
    The capillary number is used to determine which forces dominate in a specific scenario. When Ca>>1, surface forces are dominated by the viscous forces. When Ca< ...
  8. [8]
    Capillary Number - an overview | ScienceDirect Topics
    The capillary number is a dimensionless group that represents the ratio of viscous forces to the interfacial forces affecting the flow of fluid in porous media.
  9. [9]
  10. [10]
    Buckingham π Theorem - an overview | ScienceDirect Topics
    Buckingham π theorem (also known as Pi theorem) is used to determine the number of dimensional groups required to describe a phenomena.
  11. [11]
  12. [12]
    Water - Dynamic and Kinematic Viscosity at Various Temperatures ...
    The calculator below can be used to calculate the liquid water dynamic or kinematic viscosity at given temperatures. The output dynamic viscosity is given as cP ...
  13. [13]
    Surface Tension - Water in contact with Air - The Engineering ToolBox
    Surface tension of water in contact with air for temperatures ranging 0 to 100 degC (32 to 212 degF) - in imperial units (BG units) and SI units.
  14. [14]
  15. [15]
    Capillary number correlations for two-phase flow in porous media
    The analysis introduces generalized local macroscopic capillary number correlations. It eliminates shortcomings of conventional capillary number correlations.
  16. [16]
    Experimental investigations of non‐Newtonian/Newtonian liquid ...
    Mar 9, 2017 · The power-law model25 was used to fit the data (see inset in Figure ... Using a modified Capillary number to account for the non-Newtonian ...
  17. [17]
    Electro-Hydrodynamics of Emulsion Droplets: Physical Insights to ...
    Electric capillary number is defined as the ratio between the magnitude of electrical stresses ( ε 2 E 0 2 ) and capillary stresses ( γ / a ), where γ is the ...
  18. [18]
    Numerical, experimental, and theoretical investigation of bubble ...
    Jul 2, 2010 · ... (sometimes referred to as the magnetic capillary number),. Bo M = μ 0 χ H 2 d σ ,. (6). where d is the characteristic length of the bubble, ...
  19. [19]
    Investigation of interface deformation dynamics during high-Weber ...
    Capillary and viscous forces. Capillary and viscous forces may be negligible at high Weber and low Ohnesorge numbers. We assess the significance of these ...
  20. [20]
    On the Dynamic Contact Angle of Capillary-Driven Microflows in ...
    where Ca is the capillary number (Ca = μV/γ, μ being the fluid viscosity in mPa·s and g the fluid surface tension), and χ is a parameter given by χ = ln(Lma ...2.2. Solvent Prep And... · 3. Theoretical Approach · Figure 9
  21. [21]
  22. [22]
    Snap-Off during Imbibition in Porous Media: Mechanisms ... - MDPI
    Nov 17, 2023 · This implies that a higher capillary number amplifies the influence of viscous forces in the continuous phase (wetting phase), facilitating the ...
  23. [23]
    Capillary Desaturation Curve Fundamentals - OnePetro
    Apr 11, 2016 · We present a physical model to describe the capillary desaturation curve. The model balances the capillary pressure and applied viscous stresses ...Missing: seminal | Show results with:seminal
  24. [24]
    [PDF] Dimensionless Groups For Understanding Free Surface Flows ... - MIT
    Reynolds number based on a characteristic 'capillary velocity' Vcap = σ η0 (i.e. the velocity at which a viscous thread of fluid such as glycerol or pancake ...
  25. [25]
    Competition of gravity, capillary and viscous forces during drainage ...
    The capillary number Ca is the typical ratio of the viscous pressure drop at pore scale to the capillary pressure, while the “Bond number” quantifies that of ...
  26. [26]
    Bond Number | Fundamentals of Fluid Flow in Porous Media
    The Bond number is the ratio of gravity forces to capillary forces, important in vertical displacements and gravity assisted processes.
  27. [27]
    [PDF] Impact of Capillary and Bond Numbers on Relative Permeability
    ability, five different viscosity ratios are used for same capillary number, thus keeping capillary ... of wetting phase and the retention force for the wetting ...
  28. [28]
    A visualization study on two-phase gravity drainage in porous media ...
    In this paper, three dimensionless parameters, namely, the gravity number, the capillary number and the Bond number, were used to describe the effect of the ...
  29. [29]
    Proper Use of Capillary Number in Chemical Flooding
    Mar 1, 2017 · Higher capillary numbers were required to mobilize the wetting phase than the nonwetting phase residual saturation. The critical capillary ...
  30. [30]
    Numerical approach for enhanced oil recovery with surfactant flooding
    The changes of interconnected fractional flows and Capillary numbers in the reservoir, determine how and how much oil could produce. Surfactants prevent the ...
  31. [31]
    Updated Perceptions on Polymer-Based Enhanced Oil Recovery ...
    This paper is a reference for novel polymers, and their hybrid techniques, to improve polymer-based cEOR field applications under HTHS conditions in carbonates.
  32. [32]
    Role of Surfactant Flooding in Enhanced Oil Recovery - Stanford
    Nov 2, 2015 · Increasing the Capillary Number (Ca). The capillary number (Ca) is defined as the ratio of viscous to capillary forces. It is given by. Ca ...<|separator|>
  33. [33]
    Surfactant flooding: The influence of the physical properties on the ...
    Surfactants adsorb on the oil/water interface, reducing the interfacial tension (IFT) and capillary forces, which are responsible for the trapping phenomena in ...
  34. [34]
    A Critical Review of Alkaline Flooding: Mechanism, Hybrid ... - MDPI
    The capillary number also relates to oil recovery. By increasing the capillary number, the mobility of oil increases. Generally, the capillary number increases ...
  35. [35]
    Variations in Wettability and Interfacial Tension during Alkali ...
    Both effects promote an increase in the capillary number, leading to incremental oil recovery.
  36. [36]
    The Application of the Buckley-Leverett Frontal Advance Theory to ...
    For simplification, the fractional flow equation is written in the above form because the capillary forces always increase the fractional flow of the displacing ...
  37. [37]
    Displacement Stability of Water Drives in Water-Wet Connate-Water ...
    Feb 1, 1974 · We therefore consider the stability of Buckley-Leverett displacement to be better than that of the Muskat type of displacement. Capillary forces ...
  38. [38]
    [PDF] Modeling Chemical EOR Processes: Some Illustrations from Lab to ...
    Apr 19, 2013 · The Capillary Desaturation Curve (CDC) relationship,. Sor. = Sor(Nc), relates the residual oil saturation as a function of the capillary number ...
  39. [39]
    [PDF] EORI Bulletin - Enhanced Oil Recovery Institute
    May 3, 2021 · Polymer flooding in Wyoming first began in 1967 at the Skull Creek Oil Field in Weston County (Lane, 1970).
  40. [40]
    [PDF] BartlesvilleProjectOffice U.S.DEPARTMENTOFENERGY Bartlesville ...
    Alkaline flooding has always being proclaimed as a potential low cost EOR process; however, high chemical consumption has prevented successful field ...<|separator|>
  41. [41]
  42. [42]
  43. [43]
  44. [44]
    Design automation of microfluidic single and double emulsion ...
    Jan 2, 2024 · Double emulsions were generated using PDMS microfluidic devices fabricated as described above. A variety of inner and outer fluids commonly used ...
  45. [45]
    [PDF] Recent advances in microfluidic platforms for single-cell ... - UTEP
    May 17, 2019 · The PDMS SIM-Chip enabled the isolation of single cells from a heterogeneous cell mixture (whole blood samples) by applying a magnetic field ...
  46. [46]
    Capillary microfluidics for diagnostic applications - Frontiers
    This review examines the fundamental concepts of capillary-driven microfluidics, emphasizing significant progress in the design of capillary pumps and valves.
  47. [47]
    Microfluidic particle dam for direct visualization of SARS-CoV-2 ...
    Jun 3, 2022 · Here, we report a decentralized, instrument-free microfluidic device that directly visualizes SARS-CoV-2 antibody levels.<|control11|><|separator|>